Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Mitochondrial damage activates the NLRP10 inflammasome

Abstract

Upon detecting pathogens or cell stress, several NOD-like receptors (NLRs) form inflammasome complexes with the adapter ASC and caspase-1, inducing gasdermin D (GSDMD)-dependent cell death and maturation and release of IL-1β and IL-18. The triggers and activation mechanisms of several inflammasome-forming sensors are not well understood. Here we show that mitochondrial damage activates the NLRP10 inflammasome, leading to ASC speck formation and caspase-1-dependent cytokine release. While the AIM2 inflammasome can also sense mitochondrial demise by detecting mitochondrial DNA (mtDNA) in the cytosol, NLRP10 monitors mitochondrial integrity in an mtDNA-independent manner, suggesting the recognition of distinct molecular entities displayed by the damaged organelles. NLRP10 is highly expressed in differentiated human keratinocytes, in which it can also assemble an inflammasome. Our study shows that this inflammasome surveils mitochondrial integrity. These findings might also lead to a better understanding of mitochondria-linked inflammatory diseases.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: m-3M3FBS triggers mitochondrial damage and NLRP10 inflammasome activation.
Fig. 2: Mitochondrial permeability transition activates NLRP10 and AIM2 inflammasomes.
Fig. 3: Molecular characterization of NLRP10 activation.
Fig. 4: Keratinocyte differentiation drives NLRP10 expression, enabling responses to mitochondrial damage.

Similar content being viewed by others

Data availability

Representative images for all ASC speck formation quantifications are deposited in Mendeley Data using the following link: https://data.mendeley.com/datasets/42fsz64kn5/1. Source data are provided with this paper. All other data are available in the article Supplementary files or from the corresponding author upon reasonable request.

References

  1. Broz, P. & Dixit, V. M. Inflammasomes: mechanism of assembly, regulation and signalling. Nat. Rev. Immunol. 16, 407–420 (2016).

    Article  CAS  PubMed  Google Scholar 

  2. Broderick, L., de Nardo, D., Franklin, B. S., Hoffman, H. M. & Latz, E. The inflammasomes and autoinflammatory syndromes. Annu. Rev. Pathol. Mech. Dis. 10, 395–424 (2015).

    Article  CAS  Google Scholar 

  3. Mangan, M. S. J. et al. Targeting the NLRP3 inflammasome in inflammatory diseases. Nat. Rev. Drug Discov. 17, 588–606 (2018).

  4. Liston, A. & Masters, S. L. Homeostasis-altering molecular processes as mechanisms of inflammasome activation. Nat. Rev. Immunol. 17, 208–214 (2017).

  5. Muñoz-Planillo, R. et al. K+ efflux is the common trigger of NLRP3 inflammasome activation by bacterial toxins and particulate matter. Immunity 38, 1142–1153 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  6. He, Y., Hara, H. & Núñez, G. Mechanism and regulation of NLRP3 inflammasome activation. Trends Biochem. Sci. 41, 1012–1021 (2016).

  7. Shi, H. et al. NLRP3 activation and mitosis are mutually exclusive events coordinated by NEK7, a new inflammasome component. Nat. Immunol. 17, 250–258 (2016).

    Article  CAS  PubMed  Google Scholar 

  8. Schmid-Burgk, J. L. et al. A genome-wide CRISPR (clustered regularly interspaced short palindromic repeats) screen identifies NEK7 as an essential component of NLRP3 inflammasome activation. J. Biol. Chem. 291, 103–109 (2016).

    Article  CAS  PubMed  Google Scholar 

  9. Chen, J. & Chen, Z. J. PtdIns4P on dispersed trans-Golgi network mediates NLR P3 inflammasome activation. Nature 564, 71–76 (2018).

  10. Iyer, S. S. et al. Mitochondrial cardiolipin is required for Nlrp3 inflammasome activation. Immunity 39, 311–323 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Bae, Y. S. et al. Identification of a compound that directly stimulates phospholipase C activity. Mol. Pharmacol. 63, 1043–1050 (2003).

    Article  CAS  PubMed  Google Scholar 

  12. Lee, G.-S. et al. The calcium-sensing receptor regulates the NLRP3 inflammasome through Ca 21 and cAMP. Nature 492, 123–127 (2012).

  13. Rossol, M. et al. Extracellular Ca2+ is a danger signal activating the NLRP3 inflammasome through G protein-coupled calcium sensing receptors. Nat. Commun. 3, 1329 (2012).

    Article  PubMed  Google Scholar 

  14. Mariathasan, S. et al. Cryopyrin activates the inflammasome in response to toxins and ATP. Nature 440, 228–232 (2006).

  15. Fernandes-Alnemri, T., Yu, J.-W., Datta, P., Wu, J. & Alnemri, E. S. AIM2 activates the inflammasome and cell death in response to cytoplasmic DNA. Nature 458, 509–513 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Hornung, V. et al. AIM2 recognizes cytosolic dsDNA and forms a caspase-1-activating inflammasome with ASC. Nature 458, 514–518 (2009).

  17. Bauernfeind, F. G. et al. Cutting edge: NF-κB activating pattern recognition and cytokine receptors license NLRP3 inflammasome activation by regulating NLRP3 expression. J. Immunol. 183, 787–791 (2009).

    Article  CAS  PubMed  Google Scholar 

  18. Franchi, L., Eigenbrod, T. & Núñez, G. Cutting edge: TNF-α mediates sensitization to ATP and silica via the NLRP3 inflammasome in the absence of microbial stimulation. J. Immunol. 183, 792–796 (2009).

  19. Coll, R. C. et al. A small-molecule inhibitor of the NLRP3 inflammasome for the treatment of inflammatory diseases. Nat. Med. 21, 248–255 (2015).

  20. Coll, R. C. et al. MCC950 directly targets the NLRP3 ATP-hydrolysis motif for inflammasome inhibition. Nat. Chem. Biol. 15, 556–559 (2019).

    Article  CAS  PubMed  Google Scholar 

  21. Friberg, H., Ferrand-Drake, M., Bengtsson, F., Halestrap, A. P. & Wieloch, T. Cyclosporin A, but not FK 506, protects mitochondria and neurons against hypoglycemic damage and implicates the mitochondrial permeability transition in cell death. J. Neurosci. 18, 5151–5159 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Korge, P. & Weiss, J. N. Thapsigargin directly induces the mitochondrial permeability transition. Eur. J. Biochem. 265, 273–280 (1999).

    Article  CAS  PubMed  Google Scholar 

  23. Xin, M. et al. Small-molecule Bax agonists for cancer therapy. Nat. Commun. 5, 4935 (2014).

  24. Lech, M., Avila-Ferrufino, A., Skuginna, V., Susanti, H. E. & Anders, H.-J. Quantitative expression of RIG-like helicase, NOD-like receptor and inflammasome-related mRNAs in humans and mice. Int. Immunol. 22, 717–728 (2010).

  25. Lautz, K. et al. NLRP10 enhances Shigella-induced pro-inflammatory responses. Cell. Microbiol. 14, 1568–1583 (2012).

  26. Asano, M. et al. Characterization of innate and adaptive immune responses in PYNOD-deficient mice. Immunohorizons 2, 139–141 (2018).

    Google Scholar 

  27. Vacca, M. et al. NLRP10 enhances CD4+ T-cell-mediated IFNγ response via regulation of dendritic cell-derived IL-12 release. Front. Immunol. 8, 1462 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Dang, E. V., McDonald, J. G., Russell, D. W. & Cyster, J. G. Oxysterol restraint of cholesterol synthesis prevents AIM2 inflammasome activation. Cell 171, 1057–1071.e11 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Baines, C. P. et al. Loss of cyclophilin D reveals a critical role for mitochondrial permeability transition in cell death. Nature 434, 658–662 (2005).

    Article  CAS  PubMed  Google Scholar 

  30. Basso, E. et al. Properties of the permeability transition pore in mitochondria devoid of cyclophilin D. J. Biol. Chem. 280, 18558–18561 (2005).

    Article  CAS  PubMed  Google Scholar 

  31. Nakagawa, T. et al. Cyclophilin D-dependent mitochondrial permeability transition regulates some necrotic but not apoptotic cell death. Nature 434, 652–658 (2005).

    Article  CAS  PubMed  Google Scholar 

  32. Marton, J. et al. Cyclosporine a treatment inhibits Abcc6-dependent cardiac necrosis and calcification following coxsackievirus B3 infection in mice. PLoS ONE 10, e0138222 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Steinkasserer, A. et al. Mode of action of SDZ NIM 811, a nonimmunosuppressive cyclosporin A analog with activity against human immunodeficiency virus type 1 (HIV-1): interference with early and late events in HIV-1 replication J. Virol. https://doi.org/10.1128/jvi.69.2.814-824.1995 (1995).

  34. Paeshuyse, J. et al. The non-immunosuppressive cyclosporin DEBIO-025 is a potent inhibitor of hepatitis C virus replication in vitro. Hepatology 43, 761–770 (2006).

  35. Bernardi, P. et al. The mitochondrial permeability transition from in vitro artifact to disease target. FEBS J. 273, 2077–2099 (2006).

  36. Quarato, G., Llambi, F., Guy, C.S. et al. Ca2+-mediated mitochondrial inner membrane permeabilization induces cell death independently of Bax and Bak. Cell Death Differ. 29, 1318–1334 (2022).

  37. Macdonald, J. A., Wijekoon, C. P., Liao, K.-C. & Muruve, D. A. Biochemical and structural aspects of the ATP-binding domain in inflammasome-forming human NLRP proteins. IUBMB Life 65, 851–862 (2013).

  38. Duncan, J. A. et al. Cryopyrin/NALP3 binds ATP/dATP, is an ATPase, and requires ATP binding to mediate inflammatory signaling. Proc. Natl Acad Sci. USA 104, 8041–8046 (2007).

  39. Tapia-Abellán, A. et al. MCC950 closes the active conformation of NLRP3 to an inactive state. Nat. Chem. Biol. 15, 560–564 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  40. Stack, J. H. et al. IL-Converting enzyme/caspase-1 inhibitor VX-765 blocks the hypersensitive response to an inflammatory stimulus in monocytes from familial cold autoinflammatory syndrome patients. J. Immunol. 175, 2630–2634 (2005).

  41. Hoglen, N. C. et al. Characterization of IDN-6556 (3-{2-(2-tert-butyl-phenylaminooxalyl)-amino]-propionylamino}-4-oxo-5-(2,3,5, 6-tetrafluoro-phenoxy)-pentanoic acid): a liver-targeted caspase Inhibitor. J. Pharmacol. Exp. Ther. 309, 634–640 (2004).

    Article  CAS  PubMed  Google Scholar 

  42. Mirza, N., Sowa, A. S., Lautz, K. & Kufer, T. A. NLRP10 affects the stability of Abin-1 to control inflammatory responses. J. Immunol. 202, 218–227 (2019).

    Article  CAS  PubMed  Google Scholar 

  43. Damm, A., Giebeler, N., Zamek, J., Zigrino, P. & Kufer, T. A. Epidermal NLRP10 contributes to contact hypersensitivity responses in mice. Eur. J. Immunol. 46, 1959–1969 (2016).

    Article  CAS  PubMed  Google Scholar 

  44. Lachner, J., Mlitz, V., Tschachler, E. & Eckhart, L. Epidermal cornification is preceded by the expression of a keratinocyte-specific set of pyroptosis-related genes. Sci. Rep. 7, 17446 (2017).

  45. Robinson, K. S. et al. ZAKa-driven ribotoxic stress response activates the human NLRP1 inflammasome. Science 377, 328–335 (2022).

  46. Tanaka, N. et al. Eight novel susceptibility loci and putative causal variants in atopic dermatitis. J. Allergy Clin. Immunol. 148, 1293–1306 (2021).

    Article  CAS  PubMed  Google Scholar 

  47. Hirota, T. et al. Genome-wide association study identifies eight new susceptibility loci for atopic dermatitis in the Japanese population. Nat. Genet. 44, 1222–1226 (2012).

  48. Wang, Y. et al. PYNOD, a novel Apaf‐1/CED4‐like protein is an inhibitor of ASC and caspase‐1. Int. Immunol. 16, 777–786 (2004).

  49. Kinoshita, T., Wang, Y., Hasegawa, M., Imamura, R. & Suda, T. PYPAF3, a PYRIN-containing APAF-1-like protein, is a feedback regulator of caspase-1-dependent interleukin-1β secretion. J. Biol. Chem. 280, 21720–21725 (2005).

    Article  CAS  PubMed  Google Scholar 

  50. Meunier, E. & Broz, P. Evolutionary convergence and divergence in NLR function and structure. Trends Immunol. 38, 744–757 (2017).

  51. The Tabula Muris Consortium. Single-cell transcriptomics of 20 mouse organs creates a Tabula Muris. Nature 562, 367–372 (2018).

  52. Zheng, D. et al. Epithelial NLRP10 inflammasome mediates protection against intestinal autoinflammation. Nat. Immunol. (in the press).

  53. Dickson, M. A. et al. Human keratinocytes that express hTERT and also bypass a p16(INK4a)-enforced mechanism that limits life span become immortal yet retain normal growth and differentiation characteristics. Mol. Cell. Biol. https://doi.org/10.1128/mcb.20.4.1436-1447.2000 (2000).

  54. Aiyar, A., Xiang, Y., Leis, J. Site-directed mutagenesis using overlap extension PCR. in In Vitro Mutagenesis Protocols. Methods In Molecular Medicine Vol. 57 (ed Trower, M. K.) 177–191 (Humana Press, 1996).

  55. Robinson, K. S. et al. Enteroviral 3C protease activates the human NLRP1 inflammasome in airway epithelia. Science 370, eaay2002 (2020).

    Article  CAS  PubMed  Google Scholar 

  56. Jenster, L.-M. et al. P38 kinases mediate NLRP1 inflammasome activation after ribotoxic stress response and virus infection. J. Exp. Med. 220, e20220837 (2023).

  57. Hashiguchi K. & Zhang-Akiyama, Q.-M. Establishment of human cell lines lacking mitochondrial DNA. in Mitochondrial DNA: Methods and Protocols (ed Stuart, J. A.) 383–391 (Humana Press, 2009).

  58. West, A. P. et al. Mitochondrial DNA stress primes the antiviral innate immune response. Nature 520, 553–557 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Livak, K. J. & Schmittgen, T. D. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCT method. Methods 25, 402–408 (2001).

    Article  CAS  PubMed  Google Scholar 

  60. Franklin, B. S. et al. The adaptor ASC has extracellular and ‘prionoid’ activities that propagate inflammation. Nat. Immunol. 15, 727–737 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Stutz, A., Horvath, G. L., Monks, B. G. & Latz, E. ASC speck formation as a readout for inflammasome activation. in The Inflammasome. Methods in Molecular Biology Vol. 1040 (De Nardo, C., & Latz, E.) 91–101 (Humana Press, 2013). .

  62. Carpenter, A. E. et al. CellProfiler: image analysis software for identifying and quantifying cell phenotypes. Genome Biol. 7, R100 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Kamentsky, L. et al. Improved structure, function and compatibility for CellProfiler: modular high-throughput image analysis software. Bioinformatics 27, 1179–1180 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. McQuin, C. et al. CellProfiler 3.0: next-generation image processing for biology. PLoS Biol. https://doi.org/10.1371/journal.pbio.2005970 (2018).

Download references

Acknowledgements

We thank H. Beckert, G. L. Horvath, L. Maus and P. Stausberg from the Microscopy Core Facility of the Medical Faculty at the University of Bonn for providing help, services and devices funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)–Projektnummer 13123509, 388171357 and 169331223. This work was also supported by the Light Microscopy Facility, a Core Facility of the CMCB Technology Platform at TU Dresden. We thank D. Bertheloot, D. De Nardo, Y. Herzig, B. G. Monks, L. Secchim Ribeiro, G. Engels and P. Leka (all from the Institute of Innate Immunity in Bonn) for providing reagents and expertise, as well as N. Zimmermann from the Department of Dermatology, Medical Faculty Carl Gustav Carus, Technische Universität Dresden for outstanding technical assistance. This work was funded in part by the European Union’s Horizon 2020 research and innovation program under grant agreement no. 848146 (To_Aition) (to E.L.), by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy–EXC2151–390873048 (to E.L.), by the DFG SFB1454-432325352 (to E.L.), SFB1402–414786233 (to E.L. and F.I.S.), TRR237–369799452 (to E.L. and C. Günther) and GRK2168–272482170 (to E.L.). This work was further supported by the Helmholtz-Gemeinschaft, Zukunftsthema ‘Immunology and Inflammation’ (ZT-0027, to E.L.) and the National Research Foundation-Prime Minister’s office, Republic of Singapore: NRF Fellowship 2018 (NRF-NRFF11-2019-0006) (to F.L.Z.) as well as the Nanyang Technological University: NAP2018 (to F.L.Z.) and an A*STAR Career Development Award (to K.S.R.).

Author information

Authors and Affiliations

Authors

Contributions

T.P., M.S.J.M., D.W. and E.L. conceived the project. T.P. and E.L. wrote the original draft. T.P., M.B.V., M.S.J.M., K.S.R., S.R., D.W., T.A.K., F.L.Z. and E.L. reviewed and edited the paper. T.P., M.B.V., M.S.J.M., K.S.R., S.R., F.G.D., L.S., R.K., M.R., T.A.K., F.I.S., D.D.G., M.L., R.S., K.A., N.M. and L.-M.J. devised the methodology. T.P., M.B.V., D.D.G., K.S.R., S.R., C. Gottschild, K.S., F.L.Z. and E.L. carried out the investigations. T.P., M.B.V. and E.L. provided the visualization. Funding acquisition was by C. Günther, F.L.Z., F.I.S. and E.L.

Corresponding author

Correspondence to Eicke Latz.

Ethics declarations

Competing interests

E.L. is cofounder and consultant of IFM Therapeutics and Odyssey Therapeutics as well as a cofounder and board member of Dioscure Therapeutics and a Stealth Biotech. The other authors declare no competing interests.

Peer review

Peer review information

Nature Immunology thanks the anonymous reviewers for their contribution to the peer review of this work. Primary Handling Editor: N. Bernard, in collaboration with the Nature Immunology editorial team.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 m-3M3FBS activates the inflammasome independent of NLRP3.

(a) IL-1β secretion from WT iMacs, LPS-primed (200 ng/ml) and stimulated with nigericin (10 μM), poly-(dA:dT) (2 μg/ml) or m-3M3FBS (80 μM) (n = 3). (b) ASC specks per nuclei in iMacs expressing NLRP3 and ASCmCerulean, stimulated as in (a) (n = 3). (c) IL-1β secretion from WT iMacs, LPS-primed (200 ng/ml) and stimulated with nigericin (10 μM), poly-(dA:dT) (2 μg/ml) or m-3M3FBS (85 μM), in the presence of increasing doses of KCl (0, 5, 10, 25, 50, 75, 100, 125 mM) (n = 3). (d) ASC specks per nuclei in iMacs expressing NLRP3 and ASCmCerulean, stimulated as in (c), in the presence of increasing doses of KCl (0, 5, 10, 25, 50, 75, 100, 125 mM) (n = 3). (e) ASC specks per nuclei in WT iMacs expressing ASCmCerulean, with or without LPS priming (200 ng/ml) (41, 42), stimulated as in (c). While nigericin-induced ASC speck formation depends on priming, m-3M3FBS and poly-(dA:dT) trigger ASC speck formation in both LPS-primed and -unprimed cells (n = 3). (f) IL-1β secretion from WT iMacs, LPS-primed (200 ng/ml) and stimulated as in (c), in the presence or absence of the NLRP3 inhibitor CRID3 (5 μM) (43-45) (n = 3). (g) ASC specks per nuclei in iMacs expressing NLRP3 and ASCmCerulean, stimulated as in (c), in the presence or absence of the NLRP3 inhibitor CRID3 (5 μM) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. **P = 0.0068, ****P < 0.0001, ns P = 0.5730-0.9999. (h) ASC speck formation in HEK cells stably expressing ASCTagBFP and transfected with an empty vector (EV) or human NLRC4, Pyrin or AIM2, stimulated with increasing doses of m-3M3FBS (40, 55, 70, or 85 μM); normalized to the unstimulated control (n = 3). (i) ASC specks per nuclei in HEK cells expressing human NLRP10 or mCitrine and ASCTagBFP, stimulated with nigericin (10 μM), poly-(dA:dT) (2 μg/ml) or m-3M3FBS (85 μM) (n = 4). Individual data points are means of technical triplicates (a, b, e, h, i) or duplicates (c, d, f, g); error bars represent SD.

Source data

Extended Data Fig. 2 NLRP10 does not colocalize with the ER, endolysosomal or Golgi compartments.

(a–c) Representative confocal micrographs of HEK cells stably expressing NLRP10mCherry and transfected with an endoplasmic reticulum marker (DDOSTmCitrine) (a), an endolysosomal compartment marker (LAMP1mCitrine) (b), or a Golgi apparatus marker (TGOLN2mCitrine) (c), stimulated with m-3M3FBS (85 μM) or nigericin (10 μM) (n = 4).

Extended Data Fig. 3 NLRP10 colocalizes with mitochondria and nucleates ASC specks upon thapsigargin and SMBA1 treatments.

(a) Representative confocal micrographs of HEK cells stably expressing NLRP10mCherry and transfected with TOMM20mCitrine mitochondrial marker, stimulated with thapsigargin (20 μM) or SMBA1 (50 μM) (n = 4). (b) ASC specks per nuclei in HEK cells expressing NLRP10 and ASCTagBFP, stimulated with m-3M3FBS (85 μM), thapsigargin (20 μM) or increasing doses of SMBA1 (5, 10, or 50 μM) (n = 3). Individual data points (where applicable) are means of technical duplicates (b); all error bars represent SD.

Source data

Extended Data Fig. 4 Upon mitochondrial permeabilization, AIM2 but not NLRP10 activation requires mtDNA.

(a) IL-1β secretion from WT iMacs, LPS-primed (200 ng/ml) and stimulated with thapsigargin (20 μM) (n = 3) or SMBA1 (50μM) (n = 4), in the presence or absence of the NLRP3 inhibitor CRID3 (5 μM) (43-45). P values were calculated by unpaired two-tailed Student’s t-test. ns P = 0.8492–0.8931. (b) IL-1β secretion from WT or AIM2-deficient bone marrow-derived macrophages, primed with LPS (200 ng/ml) and stimulated with poly-(dA:dT) dsDNA (2 μg/ml) or SMBA1 (50 μM) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparison test. ****P < 0.0001. (c) ASC specks per nuclei in iMacs expressing NLRP3 and ASCmCerulean, stimulated with nigericin (10 μM), poly-(dA:dT) dsDNA (2 μg/ml), m-3M3FBS (85 μM), thapsigargin (20 μM) or SMBA1 (50 μM) (n = 4 for treatment with SMBA1 and 3 for other activators). (d, e) qPCR assessment of the whole-cell mtDNA content in WT iMacs (d) or HEK cells expressing NLRP10 and ASCTagBFP (e), cultured for 72–96 h with 2’,3’-dideoxycytidine (ddC) (80 μg/ml) (n = 3). (f) ASC specks per nuclei in HEK cells expressing NLRP10 and ASCTagBFP, cultured in the presence or absence of ddC (80 μg/ml), and stimulated with m-3M3FBS (85 μM) or thapsigargin (20 μM) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. ns P = 0.1337–0.9909. Individual data points are values of technical monoplicates (d and e) or means of duplicates (a, c) or triplicates (b, f); all error bars represent SD.

Source data

Extended Data Fig. 5 Differential sensitivity to cyclosporin A links mitochondrial damage to NLRP10 and AIM2 activation.

(a, b) Representative micrographs from time-lapse microscopy of HEK cells expressing NLRP10, ASCTagBFP and mCherry targeted to the mitochondrial matrix (a) or iMacs expressing NLRP3, ASCmCerulean and mCitrine targeted to the mitochondrial matrix (b) (using cytochrome c oxidase subunit 8 mitochondrial targeting sequence). Cells were treated with m-3M3FBS (85 μM) or thapsigargin (20 μM), with or without cyclosporin A (CsA; 15 μM) pre-treatment (n = 4). Images of the same field at treatment onset (0 min) or indicated time points.

Extended Data Fig. 6 NLRP10 activation by SMBA1 depends on mitochondrial permeability transition.

(a, c) ASC specks per nuclei in HEK cells expressing NLRP10 and ASCTagBFP (a) or iMacs expressing NLRP3 and ASCmCerulean (c), treated with SMBA1 (50 μM), in the presence or absence of cyclosporin A (CsA; 10 μM) (n = 3). P values were calculated by unpaired two-tailed Student’s t-test. ***P = 0.0009, ****P < 0.0001. (b, d) Representative endpoint fluorescence micrographs of HEK cells expressing NLRP10 and ASCTagBFP (b) or iMacs expressing NLRP3 and ASCmCerulean (d) with stable expression of mCitrine targeted to the mitochondrial matrix (using mitochondrial targeting sequence from cytochrome c oxidase subunit 8), stimulated with SMBA1 (50 μM), in the presence or absence of CsA (10 μM). Note that in the presence but not in the absence of CsA pre-treatment, mCitrine retains the granular distribution pattern consistent with mitochondrial localization (n = 3). Individual data points (where applicable) are means of technical duplicates; all error bars represent SD.

Source data

Extended Data Fig. 7 Debio025 and NIM811 inhibit thapsigargin- but not m-3M3FBS-induced NLRP10 activation.

(a, d) ASC specks per nuclei in HEK cells expressing NLRP10 and ASCTagBFP, stimulated with m-3M3FBS (85 μM) or thapsigargin (20 μM), in the presence or absence of Debio025 (10 μM) (a) or NIM811 (10 μM) (d) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. *P = 0.0276, **P = 0.0038, ns P = 0.0778-0.0792. (b, e) ASC specks per nuclei in iMacs expressing NLRP3 and ASCmCerulean, stimulated with nigericin (10 μM), poly-(dA:dT) dsDNA (2 μg/ml), m-3M3FBS (85 μM) or thapsigargin (20 μM), in the presence or absence of Debio025 (10 μM) (b) or NIM811 (10 μM) (e) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. ***P = 0.0002, ****P < 0.0001, ns P = 0.2813-0.9999. (c, f) IL-1β secretion from WT iMacs, primed with LPS (200 ng/ml) and stimulated as in (b, e), in the presence or absence of Debio025 (10 μM) (c) or NIM811 (10 μM) (f) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. **P = 0.0027, ****P < 0.0001, ns P = 0.1174-0.9996. Individual data points are means of technical duplicates; all error bars represent SD.

Source data

Extended Data Fig. 8 NLRP10 activation upon m-3M3FBS treatment is independent of calcium fluxes.

(a, b) ASC specks per nuclei in HEK cells expressing NLRP10 and ASCTagBFP, with or without m-3M3FBS (85 μM) treatment, in the presence or absence of BAPTA-AM (50 μM) (a) or 2-APB (50 μM) (b) in Ca2+-free medium (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. **P = 0.0019, ns P = 0.3923. (c) Representative micrographs from time-lapse microscopy of HEK cells expressing NLRP10, ASCTagBFP and mCherry targeted to the mitochondrial matrix (using cytochrome c oxidase subunit 8 mitochondrial targeting sequence), treated with ionomycin (10 μM) in the presence of CaCl2 (1 mM). Images of the same field at treatment onset (0 min) or indicated time point (30 min) (n = 4). (d) Quantification of mitochondrial fluorescence signal granularity and ASC specks per imaging field (n = 4). (e) ASC specks per nuclei in HEK cells expressing NLRP10 and ASCTagBFP, stimulated with m-3M3FBS (85 μM) or with increasing doses of ionomycin (5, 10, 15 μM) in the presence of CaCl2 (1 mM) (n = 3). (f) ASC specks per nuclei in HEK cells expressing NLRP10 and ASCTagBFP, with or without m-3M3FBS (85 μM) treatment, in the presence or absence of ionomycin (15 μM) in Ca2+-free medium (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. *P = 0.0155. Individual data points are means of technical duplicates (a, b, e, and f) or means from 4 independent experiments (d); all error bars represent SD.

Source data

Extended Data Fig. 9 Impact of emricasan on IL-1β secretion and ASC speck formation during mitochondrial rupture.

(a) IL-1β secretion from WT iMacs, primed with LPS (200 ng/ml) and stimulated with nigericin (10 μM), poly-(dA:dT) dsDNA (2 μg/ml), m-3M3FBS (85 μM) or thapsigargin (20 μM), in the presence or absence of emricasan (5 μM) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. ****P < 0.0001. (b) ASC specks per nuclei in HEK cells expressing NLRP10 and ASCTagBFP, stimulated with m-3M3FBS (85 μM) or thapsigargin (20 μM), in the presence or absence of emricasan (5 μM) (n = 3). (c) ASC specks per nuclei in iMacs expressing NLRP3 and ASCmCerulean, stimulated as in (a), in the presence or absence of emricasan (5 μM) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. ns P = 0.4112-0.9999. Individual data points are means of technical duplicates; all error bars represent SD.

Source data

Extended Data Fig. 10 NLRP10 expression in primary and immortalized keratinocytes.

(a) qPCR analysis of NLRP10 mRNA expression in primary human keratinocytes, undifferentiated or differentiated for 6 days in medium containing CaCl2 (0.1–1 mM) (n = 2). (b) Representative western blot analysis of the differentiation marker involucrin and NLRP10 in undifferentiated or differentiated immortalized N/TERT human keratinocytes (n = 3). (c) Representative western blot analysis of NLRP10 in WT, Cas9 control or NLRP10-deficient (NLRP10sg2, NLRP10sg3) immortalized N/TERT human keratinocytes (n = 3). (d) Representative widefield fluorescence microscopy analysis of caspase-1 CARD domain aggregate formation in NLRP10-overexpressing or NLRP10-deficient immortalized N/TERT human keratinocytes, expressing Casp1-CARDeGFP, with or without m-3M3FBS (85 μM) treatment (n = 1). (e) Representative immunofluorescence staining and widefield fluorescence microscopy of NLRP10-overexpressing or NLRP10-deficient immortalized N/TERT human keratinocytes (n = 1). (f, g) Representative immunofluorescence staining and confocal microscopy of primary normal human keratinocytes, with or without m-3M3FBS (85 μM) treatment. Shown as maximum intensity projection of the z-stack, from an orthogonal (f) or top (g) view (n = 3). (h) ASC specks per nuclei in HEK cells expressing NLRP10 (either WT or the R243W variant) and ASCTagBFP, stimulated with m-3M3FBS (85 μM) or thapsigargin (20 μM) (n = 3). P values were calculated by 2-way ANOVA with Šídák’s multiple comparisons test. ***P = 0.0002-0.0005. Individual data points are means of technical duplicates (a, h); all error bars represent SD.

Source data

Supplementary information

Source data

Source Data Figs. and Extended Data Figs.

Unprocessed immunoblots of all Figs. and Extended Data Figs.

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Próchnicki, T., Vasconcelos, M.B., Robinson, K.S. et al. Mitochondrial damage activates the NLRP10 inflammasome. Nat Immunol 24, 595–603 (2023). https://doi.org/10.1038/s41590-023-01451-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-023-01451-y

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing